This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Inversion Diameter and Treewidth

Yichen Wang E-mail: [email protected] Department of Mathematical Sciences, Tsinghua University, Beijing 100084, China Haozhe Wang E-mail: [email protected] Department of Mathematical Sciences, Tsinghua University, Beijing 100084, China Yuxuan Yang Correspondence Author. E-mail: [email protected] School of Science, Beijing University of Posts and Telecommunications, Beijing 100876, China Mei Lu E-mail: [email protected] Department of Mathematical Sciences, Tsinghua University, Beijing 100084, China
Abstract

In an oriented graph G\overrightarrow{G}, the inversion of a subset XX of vertices consists in reversing the orientation of all arcs with both end-vertices in XX. The inversion graph of a graph GG, denoted by (G)\mathcal{I}(G), is the graph whose vertices are orientations of GG in which two orientations G1\overrightarrow{G_{1}} and G2\overrightarrow{G_{2}} are adjacent if and only if there is an inversion XX transforming G1\overrightarrow{G_{1}} into G2\overrightarrow{G_{2}}. The inversion diameter of a graph GG is the diameter of its inversion graph (G)\mathcal{I}(G) denoted by diam((G))diam(\mathcal{I}(G)). Havet, Hörsch, and Rambaud (2024) first proved that for GG of treewidth kk, diam((G))2kdiam(\mathcal{I}(G))\leq 2k, and there are graphs of treewidth kk with inversion diameter k+2k+2. In this paper, we construct graphs of treewidth kk with inversion diameter 2k2k, which implies that the previous upper bound diam((G))2kdiam(\mathcal{I}(G))\leq 2k is tight. Moreover, for graphs with maximum degree Δ\Delta, Havet, Hörsch, and Rambaud (2024) proved diam((G))2Δ1diam(\mathcal{I}(G))\leq 2\Delta-1 and conjectured that diam((G))Δdiam(\mathcal{I}(G))\leq\Delta. We prove the conjecture when Δ=3\Delta=3 with the help of computer calculations.

Keywords: inversion diameter; orientation; treewidth.

1 Introduction

An orientation of an undirected graph is an assignment of a direction to each edge, turning the initial graph into a directed graph. Let GG be a simple graph and G1\overrightarrow{G_{1}} an orientation of GG. If XX is a vertex set of GG, the inversion of XX on G1\overrightarrow{G_{1}} is a orientation G2\overrightarrow{G_{2}} by reversing the orientation of all arcs with both ends in XX.

The concept of inversion was first introduced by Belkhechine et al. [4]. They studied the inversion number of a directed graph DD, denoted by inv(D)inv(D), which is the minimum number of inversions that transform DD into an acyclic graph. They proved, for every fixed kk, given a tournament TT, determining whether inv(T)kinv(T)\leq k is polynomial-time solvable. In contrast, Bang-Jensen et al. [3] proved that given any directed graph DD, determining whether inv(D)1inv(D)\leq 1 is NP-complete.

The maximum inversion numbers of all oriented graphs of order nn, denoted by inv(n)inv(n), has also been investigated. Aubian et al. [2] and Alon et al. [1] proved n22logninv(n)nlog(n+1)n-2\sqrt{2\log n}\leq inv(n)\leq n-\lceil\log(n+1)\rceil. Besides these results, various related questions have also been studied.

Let GG be a simple graph. The inversion is a transformation between different orientations of GG. Instead of transforming an orientation into an acyclic orientation, it is also natural to consider the transformation between two orientations. The inversion graph of GG denoted by (G)\mathcal{I}(G), is the graph whose vertices are the orientations of GG in which two orientations G1\overrightarrow{G_{1}} and G2\overrightarrow{G_{2}} are adjacent if and only if there is an inversion XX transforming G1\overrightarrow{G_{1}} into G2\overrightarrow{G_{2}}. The inversion diameter of GG is the diameter of (G)\mathcal{I}(G), denoted by diam((G))diam(\mathcal{I}(G)). It represents the maximum number of required inversions to transform an orientation of GG into another orientation of it.

Havet et al. [5] first introduced inversion diameter and studied it on various class of graphs. Let GG be a graph and let << be a total ordering on V(G)V(G). For every pair u,uu,u^{\prime} of vertices in GG, let N<u(u)={vN(u)v<u}N_{<u^{\prime}}(u)=\{v\in N(u)\mid v<u^{\prime}\} and N>u(u)={vN(u)v>u}N_{>u^{\prime}}(u)=\{v\in N(u)\mid v>u^{\prime}\}. We simply write N<(u)N_{<}(u) for N<u(u)N_{<u}(u) and N>(u)N_{>}(u) for N>u(u)N_{>u}(u). The ordering << is tt-strong if for every uV(G)u\in V(G)

  • |N<(u)|+log(|{XV(G)vN>(u),XN<u(v)}|)<t|N_{<}(u)|+\log(|\{X\subseteq V(G)\mid\exists v\in N_{>}(u),X\subseteq N_{<u}(v)\}|)<t, if N>(u)N_{>}(u)\neq\emptyset, and

  • N<(u)tN_{<}(u)\leq t otherwise.

A graph is strongly tt-degenerate if it admits a tt-strong ordering of its vertices. Havet et al. [5] showed that

Theorem 1.1 (Havet et al. [5])

Let GG be a graph and let tt be a positive integer. If GG is strong tt-degenerate, then diam((G))tdiam(\mathcal{I}(G))\leq t.

With the help of Theorem 1.1, they showed various bounds on diam((G))diam(\mathcal{I}(G)) depending on the structure of GG as following:

Theorem 1.2 (Havet et al. [5])
  1. 1.

    For every graph GG with at least one edge and maximum degree Δ\Delta, diam((G))2Δ1diam(\mathcal{I}(G))\leq 2\Delta-1.

  2. 2.

    diam((G))12diam(\mathcal{I}(G))\leq 12 for every planar graph GG.

  3. 3.

    diam((G))2kdiam(\mathcal{I}(G))\leq 2k for ever graph GG of treewidth at most kk.

Havet et al. [5] also proved that for fixed k2k\geq 2, given a graph GG, determining whether diam((G))kdiam(\mathcal{I}(G))\leq k is NP-hard. For a graph GG with maximum degree 33 (sub-cubic graph), Havet et al. [5] showed a better bound diam((G))4diam(\mathcal{I}(G))\leq 4. Moreover, they proposed a conjecture on graphs with maximum degree Δ\Delta as Conjecture 1.3.

Conjecture 1.3 (Havet et al. [5])

For every graph GG with at least one edge and maximum degree Δ\Delta, diam((G))Δdiam(\mathcal{I}(G))\leq\Delta.

The conjecture is true for Δ2\Delta\leq 2 [5]. In this paper, we prove the conjecture when Δ=3\Delta=3. Computer assistance will be used in the proof of Theorem 1.4. A pure mathematical proof is still worth studying.

Theorem 1.4

If GG is a graph of maximum degree 33, then diam((G))3diam(\mathcal{I}(G))\leq 3.

For graphs with treewidth at most kk, Havet et al. [5] showed that there are graphs of treewidth at most kk with inversion diameter k+2k+2. In this paper, we show that the upper bound diam((G))2kdiam(\mathcal{I}(G))\leq 2k for graphs of treewidth at most kk is tight by proving Theorem 1.5. It answers a question proposed by Havet et al. in [5].

Theorem 1.5

For any positive integer kk, there are graphs of treewidth kk with inversion diameter 2k2k.

This paper is organized as follows. In Section 2, we give basic definitions and notation. The proofs of Theorems 1.5 and 1.4 are given in Sections 3 and 4, respectively.

2 Preliminary

Let G=(V,E)G=(V,E) be a graph. The distance between uu and vv, denoted by dist(u,v)dist(u,v), is the number of edges in a shortest path joining uu and vv. For any vertex uV(G)u\in V(G), denote N(u)={v|uvE(G)}N(u)=\{v~{}|~{}uv\in E(G)\}. Then d(u)=|N(u)|d(u)=|N(u)| is the degree of uu. Let Δ=Δ(G)\Delta=\Delta(G) be the maximum degree of GG. We call GG kk-regular if d(u)=kd(u)=k for any uV(G)u\in V(G). Let GG be a graph and SS a vertex subset. Let G[S]G[S] denote the subgraph of GG induced by SS. For a graph GG and a vertex vv, denote GvG-v the graph induced by V(G){v}V(G)-\{v\}. For a graph GG and a induced subgraph HH, denote GHG-H the graph induced by V(G)V(H)V(G)-V(H).

A label of GG is a mapping π:E(G)𝔽2\pi:E(G)\rightarrow\mathbb{F}_{2}. A tt-dim vector assignment of GG with the label π\pi is a mapping f:V(G)𝔽2tf:V(G)\rightarrow\mathbb{F}_{2}^{t} such that π(uv)=f(u)f(v)\pi(uv)=f(u)\cdot f(v) for every edge uvE(G)uv\in E(G), where f(u)f(v)f(u)\cdot f(v) to be the scalar product of f(u)f(u) and f(v)f(v) over 𝔽2t\mathbb{F}_{2}^{t}. Usually, we use bold letter 𝐮\mathbf{u} to represent f(u)f(u). We use 𝟎\mathbf{0}  (resp. 𝟏\mathbf{1}) to represent vectors in 𝔽2t\mathbb{F}_{2}^{t} whose coordinates are all 0 (resp. 11). We say a vector 𝐮𝔽2t\mathbf{u}\in\mathbb{F}_{2}^{t} is odd (resp. even), if 𝐮𝟏\mathbf{u\cdot 1} is one  (resp. zero), i.e., 𝐮\mathbf{u} has odd (resp. even) number of 11.

The inversion diameter has a close relationship with vector assignment as following.

Proposition 2.1 ([5])

For every graph GG and every positive integer tt, the following are equivalent.

  1. 1.

    diam((G))tdiam(\mathcal{I}(G))\leq t.

  2. 2.

    For every label π\pi, there exists a tt-dim vector assignment of GG with the label π\pi.

The treewidth of a graph GG is denoted by tw(G)tw(G). There are many ways to define treewidth. Here we give a definition of treewidth from the perspective of kk-tree.

Definition 2.2

A graph GG is a k-tree if

  1. 1.

    either it is a kk-clique,

  2. 2.

    or there exists a vertex vv such that N(v)N(v) is a kk-clique, and GvG-v is a k-tree.

We say a graph is a partial kk-tree if it is a subgraph of a kk-tree. It is known that a graph GG is a partial kk-tree, if and only if the treewidth of GG is at most kk [6, 7].

Let (𝐯𝟏,,𝐯𝐤)\mathcal{L}(\mathbf{v_{1},\ldots,v_{k}}) denote the linear space spanned by 𝐯𝟏,,𝐯𝐤\mathbf{v_{1},\ldots,v_{k}}. For two vectors 𝐯\mathbf{v} and 𝐮\mathbf{u} in 𝔽2t\mathbb{F}_{2}^{t}, we write 𝐯𝐮\mathbf{v\perp u} if 𝐯𝐮=0\mathbf{v\cdot u}=0. For a vector 𝐯𝔽2t\mathbf{v}\in\mathbb{F}_{2}^{t} and a linear space 𝐔\mathbf{U} in 𝔽2t\mathbb{F}_{2}^{t}, we write 𝐯𝐔\mathbf{v\perp U} if 𝐯𝐮\mathbf{v\perp u} for any 𝐮𝐔\mathbf{u\in U}. The orthogonal complementary space of 𝐔\mathbf{U} is 𝐔={v𝐯𝐔}\mathbf{U}^{\perp}=\{\textbf{v}\mid\mathbf{v\perp U}\}. For any positive integer kk, write [k]={1,2,,k}[k]=\{1,2,\ldots,k\}.

Definition 2.3

We say a sequence of vector 𝐯𝟏,,𝐯𝐤\mathbf{v_{1},\ldots,v_{k}} are orthogonal if 𝐯𝐢𝐯𝐣\mathbf{v_{i}\perp v_{j}} for any i,j[k]i,j\in[k] and iji\neq j. We say they are self-orthogonal if 𝐯𝐢𝐯𝐣\mathbf{v_{i}\perp v_{j}} for any i,j[k]i,j\in[k], that is, they are orthogonal and every vector is even.

Definition 2.4

A linear space 𝐔\mathbf{U} is self-orthogonal if 𝐔𝐔\mathbf{U\subseteq U^{\perp}}.

Let 𝐔\mathbf{U} be a self-orthogonal linear space. Then 𝐔\mathbf{U} is orthogonal and every vector in 𝐔\mathbf{U} is even. It is easy to verify that 𝐔\mathbf{U} is self-orthogonal if and only if it has self-orthogonal base vectors.

For a linear space 𝐔\mathbf{U} and a vector 𝐯\mathbf{v}, denote 𝐯+𝐔\mathbf{v+U} by the set {𝐯+𝐮𝐮𝐔}\{\mathbf{v+u}\mid\mathbf{u\in U}\} and denote (𝐔,𝐯)\mathcal{L}(\mathbf{U,v}) the space spanned by 𝐯\mathbf{v} and a basis of 𝐔\mathbf{U}, that is the summation space of 𝐔\mathbf{U} and (𝐯)\mathcal{L}(\mathbf{v}).

3 Proof of Theorem 1.5

For k1k\geq 1, we define a sequence of graph Gi(k)G_{i}^{(k)} with a fixed label πi(k)\pi_{i}^{(k)}. First, let G0(k)G_{0}^{(k)} be a kk-clique with an arbitrarily label π0(k)\pi_{0}^{(k)}. Then, we recursively construct Gi(k)G_{i}^{(k)} as following:

(i) for each kk-clique with vertices v1,,vkv_{1},\ldots,v_{k} in Gi1(k)G_{i-1}^{(k)} and each 𝐱=(x1,,xk)T𝔽2k\mathbf{x}=(x_{1},\ldots,x_{k})^{T}\in\mathbb{F}_{2}^{k}, we add a new vertex uu such that uvjE(Gi(k))uv_{j}\in E(G_{i}^{(k)}) and πi(k)(uvj)=xj\pi_{i}^{(k)}(uv_{j})=x_{j}, for all 1jk1\leq j\leq k ;

(ii) πi(k)|Gi1(k)=πi1(k)\pi^{(k)}_{i}|_{G^{(k)}_{i-1}}=\pi^{(k)}_{i-1}.

Since |𝔽2k|=2k|\mathbb{F}_{2}^{k}|=2^{k}, we add 2k2^{k} new vertices for each kk-clique in Gi1(k)G_{i-1}^{(k)}. By Definition 2.2, Gm(k)G_{m}^{(k)} is a kk-tree for any mm, that is, of treewidth at most kk. Since πn(k)|Gm(k)=πm(k)\pi^{(k)}_{n}|_{G^{(k)}_{m}}=\pi^{(k)}_{m} when n>mn>m, we may use π(k)\pi^{(k)} to denote the label of Gm(k)G^{(k)}_{m} for any mm. For any vertex vV(Gm(k))v\in V(G^{(k)}_{m}) with m1m\geq 1, there exists an unique nn such that vV(Gn(k))V(Gn1(k))v\in V(G^{(k)}_{n})-V(G^{(k)}_{n-1}). We say nn is the level of vv, denoted by l(v)=nl(v)=n. For a vertex set SGm(k)S\subseteq G^{(k)}_{m} with m1m\geq 1, the level of SS is defined to be the maximum level of vertices in SS denoted by l(S)l(S), that is, l(S)=maxvS{l(v)}l(S)=\max\limits_{v\in S}\{l(v)\}. Clearly, if vv is a vertex in Gm(k)G_{m}^{(k)}, then l(v)ml(v)\leq m. Similarly, if CC is a vertex set in Gm(k)G_{m}^{(k)}, then l(C)ml(C)\leq m.

Note that if HH is a subgraph of GG, then diam((H))diam((G))diam(\mathcal{I}(H))\leq diam(\mathcal{I}(G)). So (diam((Gm(k))))m0(diam(\mathcal{I}(G_{m}^{(k)})))_{m\geq 0} is an increasing sequence with upper bound 2k2k by Theorem 1.2.

Let λ(k)=limm+diam((Gm(k)))\lambda^{(k)}=\lim\limits_{m\rightarrow+\infty}diam(\mathcal{I}(G_{m}^{(k)})). Then λ(k)2k\lambda^{(k)}\leq 2k. We will show that λ(k)=2k\lambda^{(k)}=2k and then Gm(k)G_{m}^{(k)} is of inversion diameter 2k2k when mm is sufficiently large.

Next we suppose λ(k)2k1\lambda^{(k)}\leq 2k-1. Then for any mm, Gm(k)G_{m}^{(k)} has a (2k1)(2k-1)-dim vector assignment with the label π(k)\pi^{(k)} by Proposition 2.1. Thus for each vV(Gm(k))v\in V(G_{m}^{(k)}), there is a vector 𝐯𝔽22k1\mathbf{v}\in\mathbb{F}_{2}^{2k-1} corresponding to it. The following lemmas show the properties of the vectors assigned to kk-cliques in Gm(k)G^{(k)}_{m}.

Lemma 3.1

If there is a kk-clique of level mm with vertices v1,,vkv_{1},\ldots,v_{k} in Gm+1(k)G^{(k)}_{m+1}, then 𝐯𝟏,,𝐯𝐤\mathbf{v_{1}},\ldots,\mathbf{v_{k}} are linear independent.

Proof. Otherwise, without loss of generality, assume 𝐯𝟏=i=2kci𝐯𝐢\mathbf{v_{1}}=\sum_{i=2}^{k}c_{i}\mathbf{v_{i}} where ci𝔽2c_{i}\in\mathbb{F}_{2} for all 2ik2\leq i\leq k. By the construction, there exists a vertex uV(Gm+1(k))u\in V(G^{(k)}_{m+1}) connecting each viv_{i} with π(k)(uvi)=1\pi^{(k)}(uv_{i})=1 for all 2ik2\leq i\leq k and π(k)(uv1)=i=2kci+1\pi^{(k)}(uv_{1})=\sum_{i=2}^{k}c_{i}+1. Therefore,

i=2kci+1\displaystyle\sum_{i=2}^{k}c_{i}+1 =π(k)(uv1)=𝐮𝐯𝟏=𝐮i=2kci𝐯𝐢=i=2kci𝐮𝐯𝐢=i=2kciπ(k)(uvi)=i=2kci,\displaystyle=\pi^{(k)}(uv_{1})=\mathbf{u\cdot v_{1}}=\mathbf{u}\cdot\sum_{i=2}^{k}c_{i}\mathbf{v_{i}}=\sum_{i=2}^{k}c_{i}\mathbf{u\cdot v_{i}}=\sum_{i=2}^{k}c_{i}\pi^{(k)}(uv_{i})=\sum_{i=2}^{k}c_{i},

a contradiction. \square

Lemma 3.2

If there is a kk-clique in Gm+2(k)G^{(k)}_{m+2} of level mm with vertices v1,,vkv_{1},\ldots,v_{k}, and uu is a vertex of level m+1m+1 connecting all (vi)1ik{(v_{i})}_{1\leq i\leq k}, then either 𝐯𝟏,,𝐯𝐤,𝐮\mathbf{v_{1},\ldots,v_{k},u} are linear independent, or 𝐮=𝐢=𝟏𝐤𝐯𝐢\mathbf{u=\sum_{i=1}^{k}v_{i}}.

Proof. Firstly, by Lemma 3.1, 𝐯𝟏,,𝐯𝐤\mathbf{v_{1},\ldots,v_{k}} are linear independent. Note that for any 1jk1\leq j\leq k, v1,,vj1,vj+1,,vk,uv_{1},\ldots,v_{j-1},v_{j+1},\ldots,v_{k},u is also a kk-clique of level m+1m+1 in Gm+2(k)G^{(k)}_{m+2}. Then by Lemma 3.1, for any 1jk1\leq j\leq k, 𝐯𝟏,,𝐯𝐣𝟏,𝐯𝐣+𝟏,,𝐯𝐤,𝐮\mathbf{v_{1}},\ldots,\mathbf{v_{j-1}},\mathbf{v_{j+1}},\ldots,\mathbf{v_{k}},\mathbf{u} are linear independent.  Assume 𝐮=i=1kci𝐯𝐢\mathbf{u}=\sum_{i=1}^{k}c_{i}\mathbf{v_{i}} where ci𝔽2c_{i}\in\mathbb{F}_{2} for all 1ik1\leq i\leq k. If cj=0c_{j}=0 for some jj, then it contradicts that 𝐯𝟏,,𝐯𝐣𝟏,𝐯𝐣+𝟏,,𝐯𝐤,𝐮\mathbf{v_{1}},\ldots,\mathbf{v_{j-1}},\mathbf{v_{j+1}},\ldots,\mathbf{v_{k}},\mathbf{u} are linear independent. Therefore, 𝐮=𝐢=𝟏𝐤𝐯𝐢\mathbf{u=\sum_{i=1}^{k}v_{i}}. \square

Lemma 3.3

Let v1,,vkv_{1},\ldots,v_{k} be vertices of a kk-clique of level mm in Gm+2(k)G^{(k)}_{m+2} and 𝐀=(𝐯𝟏,,𝐯𝐤)T\mathbf{A}=(\mathbf{v_{1}},\ldots,\mathbf{v_{k}})^{T}. Then for any 𝐛𝔽2k\mathbf{b}\in\mathbb{F}_{2}^{k}, 𝐀𝐱=𝐛\mathbf{Ax=b} has a solution 𝐲\mathbf{y} such that either 𝐯𝟏,,𝐯𝐤,𝐲\mathbf{v_{1}},\ldots,\mathbf{v_{k}},\mathbf{y} are linear independent, or 𝐲=i=1k𝐯𝐢\mathbf{y}=\sum_{i=1}^{k}\mathbf{v_{i}}.

Proof. Let 𝐛=(b1,,bk)T\mathbf{b}=(b_{1},\ldots,b_{k})^{T}. By the construction, there exists a vertex uV(Gm+1(k))u\in V(G^{(k)}_{m+1}) of level m+1m+1 connecting (vi)1ik{(v_{i})}_{1\leq i\leq k} such that π(k)(uvi)=bi\pi^{(k)}(uv_{i})=b_{i} for all 1ik1\leq i\leq k. Then we have 𝐀𝐮=𝐛\mathbf{Au=b}. By Lemma 3.2, either 𝐯𝟏,,𝐯𝐤,𝐮\mathbf{v_{1}},\ldots,\mathbf{v_{k}},\mathbf{u} are linear independent, or 𝐮=i=1k𝐯𝐢\mathbf{u}=\sum_{i=1}^{k}\mathbf{v_{i}}. \square

The above lemmas actually work for arbitrary λ(k)\lambda^{(k)}. The following lemmas need the assumption λ(k)2k1\lambda^{(k)}\leq 2k-1. If 𝐛=𝟎\mathbf{b=0}, we have a strong conclusion.

Lemma 3.4

Let v1,,vkv_{1},\ldots,v_{k} be vertices of a kk-clique of level mm in Gm+2(k)G^{(k)}_{m+2} and 𝐀=(𝐯𝟏,,𝐯𝐤)T\mathbf{A}=(\mathbf{v_{1}},\ldots,\mathbf{v_{k}})^{T}. Then 𝐀𝐱=𝟎\mathbf{Ax=0} has a solution 𝐲\mathbf{y} such that 𝐯𝟏,,𝐯𝐤,𝐲\mathbf{v_{1}},\ldots,\mathbf{v_{k}},\mathbf{y} are linear independent.

Proof. We prove it by contradiction. By Lemma 3.1, 𝐯𝟏,,𝐯𝐤\mathbf{v_{1},\ldots,v_{k}} are linear independent. Let 𝐔\mathbf{U} be the solution space of 𝐀𝐱=𝟎\mathbf{Ax=0}. Suppose 𝐔\mathbf{U} is a subspace of (𝐯𝟏,,𝐯𝐤)\mathcal{L}(\mathbf{v_{1}},\ldots,\mathbf{v_{k}}). Since 𝐀\mathbf{A} is a k×(2k1)k\times(2k-1) matrix, dim(𝐔)=(2k1)k=k1dim(\mathbf{U})=(2k-1)-k=k-1. By letting 𝐛=𝟎\mathbf{b=0} in Lemma 3.3, we have i=1k𝐯𝐢𝐔\sum_{i=1}^{k}\mathbf{v_{i}}\in\mathbf{U}.

For each j[k]j\in[k], the solution set of 𝐀𝐱=𝐀𝐯𝐣\mathbf{Ax=Av_{j}} is in 𝐯𝐣+𝐔(𝐯𝟏,,𝐯𝐤)\mathbf{v_{j}+U}\subseteq\mathcal{L}(\mathbf{v_{1}},\ldots,\mathbf{v_{k}}). By Lemma 3.3, i=1k𝐯𝐢𝐯𝐣+𝐔\sum_{i=1}^{k}\mathbf{v_{i}}\in\mathbf{v_{j}+U}. Therefore, 𝐯𝐣𝐔\mathbf{v_{j}\in U} for any 1jk1\leq j\leq k, which contradicts that dim(𝐔)=k1dim(\mathbf{U})=k-1 because 𝐯𝟏,,𝐯𝐤\mathbf{v_{1}},\ldots,\mathbf{v_{k}} are linear independent. \square

Definition 3.5

Let CC be a pp-clique of Gm(k)G^{(k)}_{m} for some mm. CC is called a bad clique if dim(VCVC)p1dim(\textbf{V}_{C}\cap\textbf{V}_{C}^{\perp})\geq p-1, where VC=({𝐯vC})\textbf{V}_{C}=\mathcal{L}(\{\mathbf{v}\mid v\in C\}) and p1p\geq 1.

Note that a single vertex is always a bad 11-clique. If λ(k)2k1\lambda^{(k)}\leq 2k-1, “large” bad clique will finally cause contradictions. The following lemma is the main part of our proof which states that we can find “large” bad clique when mm is sufficiently large.

Lemma 3.6

If there exists a bad pp-clique of level mm in Gm+k+2(k)G^{(k)}_{m+k+2} with p<kp<k, then there exists a bad clique in Gm+k+2(k)G^{(k)}_{m+k+2} of size at least p+1p+1.

Proof. We prove it by contradiction. Suppose the pp-clique C1C_{1} with vertices v1,v2,,vpv_{1},v_{2},\ldots,v_{p} of level mm is the largest bad clique in Gm+k+2(k)G^{(k)}_{m+k+2}, where p<kp<k. Then dim(VC1)=pdim(\textbf{V}_{C_{1}})=p by Lemma 3.1. Let U=VC1VC1\textbf{U}=\textbf{V}_{C_{1}}\cap\textbf{V}_{C_{1}}^{\perp}. Then dim(U)p1dim(\textbf{U})\geq p-1 by Definition 3.5. For ϕ1,ϕ2U\phi_{1},\phi_{2}\in\textbf{U}, we have ϕ1ϕ2\phi_{1}\perp\phi_{2} which means U is self-orthogonal.

We first show that dim(U)=p1dim(\textbf{U})=p-1. Suppose dim(U)=pdim(\textbf{U})=p. Then U=VC1VC1\textbf{U}=\textbf{V}_{C_{1}}\subseteq\textbf{V}_{C_{1}}^{\perp}. Since p<kp<k, by the construction of Gm+k+2(k)G^{(k)}_{m+k+2}, there exists a vertex uu of level m+1m+1 such that uviE(Gm+k+2(k))uv_{i}\in E(G^{(k)}_{m+k+2}) and π(k)(uvi)=0\pi^{(k)}(uv_{i})=0 for each i[p]i\in[p]. Let C2C1{u}C_{2}\coloneqq{C_{1}}\cup\{u\}. By Lemma 3.1, we have 𝐮,𝐯𝟏,,𝐯𝐩\mathbf{u,v_{1},\ldots,v_{p}} are linear independent. By π(k)(uvi)=0\pi^{(k)}(uv_{i})=0 for each i[p]i\in[p], we have that 𝐮𝐕𝐂𝟏\mathbf{u\perp V_{C_{1}}}. Thus VC1VC2VC2\textbf{V}_{C_{1}}\subseteq\textbf{V}_{C_{2}}\cap\textbf{V}_{C_{2}}^{\perp} which implies dim(VC2VC2)pdim(\textbf{V}_{C_{2}}\cap\textbf{V}_{C_{2}}^{\perp})\geq p. Hence C2C_{2} is a bad (p+1)(p+1)-clique, a contradiction with the maximality of C1{C_{1}}. So dim(U)=p1dim(\textbf{U})=p-1. In fact, we conclude that U is a self-orthogonal (p1)(p-1)-dim subspace of 𝔽22k1\mathbb{F}_{2}^{2k-1} and then each vector in U is even.

Since dim(U)=p1dim(\textbf{U})=p-1, there is vi{v1,v2,,vp}\textbf{v}_{i}\in\{\textbf{v}_{1},\textbf{v}_{2},\dots,\textbf{v}_{p}\}, say i=1i=1, such that v1U\textbf{v}_{1}\notin\textbf{U}. Then (U,v1)=VC1\mathcal{L}(\textbf{U},\textbf{v}_{1})=\textbf{V}_{C_{1}}. If 𝐯𝟏\mathbf{v_{1}} is even, then 𝐯𝟏(U,v1)\mathbf{v_{1}}\perp\mathcal{L}(\textbf{U},\textbf{v}_{1}), which contradicts with 𝐯𝟏U\mathbf{v_{1}}\notin\textbf{U}. Thus we have v1\textbf{v}_{1} is odd.

Claim 3.7

If uu is a vertex in Gm+k(k)G^{(k)}_{m+k} such that uviE(Gm+k(k))uv_{i}\in E(G^{(k)}_{m+k}) and π(k)(uvi)=0\pi^{(k)}(uv_{i})=0 for each i[p]i\in[p], then u is odd.

Proof of Claim 3.7. Suppose u is even. Then u(u,VC1)\textbf{u}\perp\mathcal{L}(\textbf{u},\textbf{V}_{C_{1}}). We have uVC1{u}VC1{u}\textbf{u}\in\textbf{V}_{C_{1}\cup\{u\}}\cap\textbf{V}_{C_{1}\cup\{u\}}^{\perp} and UVC1{u}VC1{u}\textbf{U}\subseteq\textbf{V}_{C_{1}\cup\{u\}}\cap\textbf{V}_{C_{1}\cup\{u\}}^{\perp}. From Lemma 3.1 and dim(U)=p1dim(\textbf{U})=p-1, we know uU\textbf{u}\notin\textbf{U}. Then dim(VC1{u}VC1{u})pdim(\textbf{V}_{C_{1}\cup\{u\}}\cap\textbf{V}_{C_{1}\cup\{u\}}^{\perp})\geq p. Thus C1{u}C_{1}\cup\{u\} is a bad (p+1)(p+1)-clique, which contradicts with the maximality of C1C_{1}.  

Fix a vertex v0v_{0} of level m+1m+1 such that v0viE(Gm+k(k))v_{0}v_{i}\in E(G^{(k)}_{m+k}) and π(k)(v0vi)=0\pi^{(k)}(v_{0}v_{i})=0 for each i[p]i\in[p]. Then 𝐯0𝐕𝐂𝟏\mathbf{v}_{0}\perp\mathbf{V_{C_{1}}}. By the construction of Gm+k(k)G^{(k)}_{m+k}, there exists a set of vertices C2={w1,w2,,wkp}V(Gm+k(k))C_{2}=\{w_{1},w_{2},\dots,w_{k-p}\}\subseteq V(G^{(k)}_{m+k}) satisfying the following:

  1. 1.

    {v0,v1,,vp,w1,w2,,wkp}\{v_{0},v_{1},\ldots,v_{p},w_{1},w_{2},\ldots,w_{k-p}\} is a (k+1)(k+1)-clique;

  2. 2.

    π(k)(wiwj)=0\pi^{(k)}(w_{i}w_{j})=0, for each i,j[kp],iji,j\in[k-p],i\neq j;

  3. 3.

    and π(k)(viwj)=vi(v1+v0)\pi^{(k)}(v_{i}w_{j})=\textbf{v}_{i}\cdot(\textbf{v}_{1}+\textbf{v}_{0}), for each i[0,p],j[kp]i\in[0,p],j\in[k-p].

Claim 3.8

wj\textbf{w}_{j} is even for each j[kp]j\in[k-p].

Proof of Claim 3.8. Suppose there is j[kp]j\in[k-p] such that wj\textbf{w}_{j} is odd. Let βwj+v1\beta\coloneqq\textbf{w}_{j}+\textbf{v}_{1}. Recall that v1\textbf{v}_{1} is odd. Then

βwj=wjwj+v1wj=wjwj+v1(v1+v0)=0.\beta\cdot\textbf{w}_{j}=\textbf{w}_{j}\cdot\textbf{w}_{j}+\textbf{v}_{1}\cdot\textbf{w}_{j}=\textbf{w}_{j}\cdot\textbf{w}_{j}+\textbf{v}_{1}\cdot(\textbf{v}_{1}+\textbf{v}_{0})=0.

On the other hand, for any viC1v_{i}\in C_{1},

βvi=wjvi+v1vi=viv1+viv0+v1vi=0.\beta\cdot\textbf{v}_{i}=\textbf{w}_{j}\cdot\textbf{v}_{i}+\textbf{v}_{1}\cdot\textbf{v}_{i}=\textbf{v}_{i}\cdot\textbf{v}_{1}+\textbf{v}_{i}\cdot\textbf{v}_{0}+\textbf{v}_{1}\cdot\textbf{v}_{i}=0.

Hence βVC1{wj}\beta\perp\textbf{V}_{C_{1}\cup\{w_{j}\}}. We have βVC1{wj}VC1{wj}\beta\in\textbf{V}_{C_{1}\cup\{w_{j}\}}\cap\textbf{V}_{C_{1}\cup\{w_{j}\}}^{\perp} and UVC1{wj}VC1{wj}\textbf{U}\subseteq\textbf{V}_{C_{1}\cup\{w_{j}\}}\cap\textbf{V}_{C_{1}\cup\{w_{j}\}}^{\perp}. From Lemma 3.1 and dim(U)=p1dim(\textbf{U})=p-1, we know wjU\textbf{w}_{j}\notin\textbf{U}. Then dim(VC1{wj}VC1{wk})pdim(\textbf{V}_{C_{1}\cup\{w_{j}\}}\cap\textbf{V}_{C_{1}\cup\{w_{k}\}}^{\perp})\geq p which implies C1{wj}C_{1}\cup\{w_{j}\} is a bad (p+1)(p+1)-clique, a contradiction with the maximality of C1C_{1}.  

v0\textbf{v}_{0} v1\textbf{v}_{1} vi\textbf{v}_{i} wj1\textbf{w}_{j_{1}} wj2\textbf{w}_{j_{2}} βj1\beta_{j_{1}} βj2\beta_{j_{2}}
v0\textbf{v}_{0} 1 0 0 1 1 0 0
v1\textbf{v}_{1} 0 1 v1vi\textbf{v}_{1}\cdot\textbf{v}_{i} 1 1 0 0
vi\textbf{v}_{i} 0 v1vi\textbf{v}_{1}\cdot\textbf{v}_{i} vivi\textbf{v}_{i}\cdot\textbf{v}_{i} v1vi\textbf{v}_{1}\cdot\textbf{v}_{i} v1vi\textbf{v}_{1}\cdot\textbf{v}_{i} 0 0
wj1\textbf{w}_{j_{1}} 1 1 v1vi\textbf{v}_{1}\cdot\textbf{v}_{i} 0 0 0 0
wj2\textbf{w}_{j_{2}} 1 1 v1vi\textbf{v}_{1}\cdot\textbf{v}_{i} 0 0 0 0
βj1\beta_{j_{1}} 0 0 0 0 0 0 0
βj2\beta_{j_{2}} 0 0 0 0 0 0 0
Table 1: The inner products for i[p]i\in[p] and j1,j2[kp]j_{1},j_{2}\in[k-p]

Now we complete the proof of Lemma 3.6. For each j[kp]j\in[k-p], let βj=v0+v1+wj\beta_{j}=\textbf{v}_{0}+\textbf{v}_{1}+\textbf{w}_{j}. By Claim 3.7, v0\textbf{v}_{0} is odd. By Claim 3.8, wj\textbf{w}_{j} is even for each j[kp]j\in[k-p]. It is not difficult to show that βjv0\beta_{j}\perp\textbf{v}_{0}, βjVC1\beta_{j}\perp\textbf{V}_{C_{1}} and βjVC2\beta_{j}\perp\textbf{V}_{C_{2}}. Check Table 1 for the inner products between the vectors that we are working on. Let C3=C1C2C_{3}=C_{1}\cup C_{2} and W=VC3=VC1+VC2W=\textbf{V}_{C_{3}}=\textbf{V}_{C_{1}}+\textbf{V}_{C_{2}}. Then dim(W)=kdim(W)=k from Lemma 3.1. Since W𝔽22k1W\subseteq\mathbb{F}_{2}^{2k-1}, we have dim(W)=k1dim(W^{\perp})=k-1. Let W=(U,β1,,βkp)W^{\prime}=\mathcal{L}(\textbf{U},\beta_{1},\dots,\beta_{k-p}). Since UW\textbf{U}\perp W and βjW\beta_{j}\perp W for each j[kp]j\in[k-p], we have WWW^{\prime}\subseteq W^{\perp}. Note that VC1,VC2(v0,v1,W)\textbf{V}_{C_{1}},\textbf{V}_{C_{2}}\subseteq\mathcal{L}(\textbf{v}_{0},\textbf{v}_{1},W^{\prime}). We have W(v0,v1,W)W\subseteq\mathcal{L}(\textbf{v}_{0},\textbf{v}_{1},W^{\prime}) which implies dim(W)k2dim(W^{\prime})\geq k-2. If v0W\textbf{v}_{0}\notin W, then dim(W)k1dim(W^{\prime})\geq k-1 which implies W=WW^{\perp}=W^{\prime}. Since v0W\textbf{v}_{0}\perp W^{\prime}, we have v0W\textbf{v}_{0}\perp W^{\perp}, a contradiction with v0W\textbf{v}_{0}\notin W. Hence v0W\textbf{v}_{0}\in W. By Lemma 3.2, we have v0++vp+w1++wkp=0\textbf{v}_{0}+\dots+\textbf{v}_{p}+\textbf{w}_{1}+\dots+\textbf{w}_{k-p}=0.

Since v0W\textbf{v}_{0}\in W, we have βjW\beta_{j}\in W for each jj which implies WWW^{\prime}\subseteq W. So WWWW^{\prime}\subseteq W\cap W^{\perp}. If dim(W)k1dim(W^{\prime})\geq k-1, then C3C_{3} is a bad kk-clique, a contradiction with the maximality of C1C_{1}. Hence dim(W)=k2dim(W^{\prime})=k-2. Let αW\W\alpha\in W^{\perp}\backslash W^{\prime} such that W=(W,α)W^{\perp}=\mathcal{L}(W^{\prime},\alpha). By the construction of Gm+k(k)G^{(k)}_{m+k}, there exists a vertex xx connecting to all vertices of C3C_{3} such that π(k)(xy)=0\pi^{(k)}(xy)=0 for each yC3y\in C_{3}. Then xW\textbf{x}\in W^{\perp}. From Claim 3.7, x is odd. Since U is a self-orthogonal subspace and βiβj\beta_{i}\bot\beta_{j} for any i,j[kp]i,j\in[k-p] (see Table 1), we have that the vectors in WW^{\prime} are all even. By xW=(W,α)\textbf{x}\in W^{\perp}=\mathcal{L}(W^{\prime},\alpha) and x being odd, we have α\alpha is odd.

Let C={v0,,vp,w1,,wkp1}C^{*}=\{v_{0},\ldots,v_{p},w_{1},\ldots,w_{k-p-1}\} (if p=k1p=k-1, then let C={v0,,vp}C^{*}=\{v_{0},\ldots,v_{p}\}). Then there is xx^{*} connecting to all vertices of CC^{*} such that π(k)(xy)=v0y\pi^{(k)}(x^{*}y)=\textbf{v}_{0}\cdot\textbf{y} for each yCy\in C^{*}. Since v0++vp+w1++wkp=0\textbf{v}_{0}+\dots+\textbf{v}_{p}+\textbf{w}_{1}+\dots+\textbf{w}_{k-p}=0, by Lemma 3.1, W=VCW=\textbf{V}_{C^{*}}. Then xv0+W\textbf{x}^{*}\in\textbf{v}_{0}+W^{\perp}. Since π(k)(xvi)=v0vi=0\pi^{(k)}(x^{*}v_{i})=\textbf{v}_{0}\cdot\textbf{v}_{i}=0 for each i[k]i\in[k], from Claim 3.7, x\textbf{x}^{*} is odd. Note that xv0+(α,W)\textbf{x}^{*}\in\textbf{v}_{0}+\mathcal{L}(\alpha,W^{\prime}). Since all vectors in WW^{\prime} are even and v0,α\textbf{v}_{0},\alpha are odd, we have xv0+WW\textbf{x}^{*}\in\textbf{v}_{0}+W^{\prime}\subseteq W. From Lemma 3.2, x=v0++vp+w1++wkp1=wkp\textbf{x}^{*}=\textbf{v}_{0}+\dots+\textbf{v}_{p}+\textbf{w}_{1}+\dots+\textbf{w}_{k-p-1}=\textbf{w}_{k-p}. Since x𝐯𝟏=01=wkp𝐯𝟏\textbf{x}^{*}\cdot\mathbf{v_{1}}=0\neq 1=\textbf{w}_{k-p}\cdot\mathbf{v_{1}} , we derive a contradiction. \square

With the help of those lemmas, we can derive a contradiction when λ(k)2k1\lambda^{(k)}\leq 2k-1 and hence, λ(k)=2k\lambda^{(k)}=2k.

Theorem 3.9

λ(k)=2k\lambda^{(k)}=2k.

Proof. Suppose λ(k)2k1\lambda^{(k)}\leq 2k-1. By Lemma 3.6, the largest bad clique C0C_{0} in Gk(k+2)(k)G^{(k)}_{k(k+2)} is of size kk. Then dim(VC0VC0)k1dim(\textbf{V}_{C_{0}}^{\perp}\cap\textbf{V}_{C_{0}})\geq k-1. By Lemma 3.1, dim(VC0)=kdim(\textbf{V}_{C_{0}})=k and then dim(VC0)=k1dim(\textbf{V}_{C_{0}}^{\perp})=k-1 by VC0𝔽22k1\textbf{V}_{C_{0}}\subseteq\mathbb{F}_{2}^{2k-1}. We have VC0VC0\textbf{V}_{C_{0}}^{\perp}\subseteq\textbf{V}_{C_{0}} by checking the dimensions. Then we can derive a contradiction by Lemma 3.4. \square

Now we can give the proof of our main Theorem.

Proof of Theorem 1.5. For any k1k\geq 1, we have λ(k)=2k\lambda^{(k)}=2k by Theorem 3.9. Then there exists a MkM_{k} such that for every mMkm\geq M_{k}, diam((Gm(k)))=2kdiam(\mathcal{I}(G^{(k)}_{m}))=2k. Thus for all mMkm\geq M_{k}, Gm(k)G^{(k)}_{m} are desired graphs of treewidth at most kk and inversion diameter 2k2k. \square

Refer to caption
Figure 1: An example of outer-planar graph with labeled edges of inversion diameter 44 verified by computer.

Note that any outer-planar graph is of treewidth 22 and hence has inversion diameter at most 44 by Lemma 1.2. We construct an outer-planar graph of inversion diameter 44 verified by computer as Figure 1. The codes are available on GitHub. Therefore, the upper bound diam((G))4diam(\mathcal{I}(G))\leq 4 for any outer-planar graph GG is tight.

4 Proof of Theorem 1.4

In this section, we intend to give the proof of Theorem 1.4.

Let GG be a graph. We say GG is 33-critical if diam((G))>3diam(\mathcal{I}(G))>3 and for any proper subgraph GG^{\prime}, diam((G))3diam(\mathcal{I}(G^{\prime}))\leq 3. Clearly, a 33-critical graph is connected. If GG is 33-critical, by Proposition 2.1, there exists a label π\pi such that there is no 33-dim vector assignment of GG with π\pi. We call such π\pi a bad label.

Let GG be a 33-critical graph with a bad label π\pi and HH an induced subgraph of GG. Denote NG(H)={vV(G)V(H)uV(H),uvE(G)}N_{G}(H)=\{v\in V(G)-V(H)\mid\exists u\in V(H),uv\in E(G)\}. By the definition of 33-critical graph, GHG-H admits a 33-dim vector assignment f:V(GH)𝔽23f:V(G-H)\rightarrow\mathbb{F}_{2}^{3} with π|GH\pi|_{G-H}. For a vertex vNG(H)v\in N_{G}(H), define 𝒜f(v)={𝐯𝔽23𝐯f(u)=π(uv), for any uvE(GH)}\mathcal{A}_{f}(v)=\{\mathbf{v}\in\mathbb{F}_{2}^{3}\mid\mathbf{v\cdot}f(u)=\pi(uv),\mbox{~{}for any~{}}uv\in E(G-H)\}. Note that f(v)𝒜f(v)f(v)\in\mathcal{A}_{f}(v).

Let HH be a subgraph of GG. We say (f(v))vNG(H)(\mathcal{B}_{f}(v))_{v\in N_{G}(H)} is an available boundary family if we can assign each vertex vNG(H)v\in N_{G}(H) a set f(v)\mathcal{B}_{f}(v) such that:

  1. 1.

    f(v)f(v)𝒜f(v)f(v)\in\mathcal{B}_{f}(v)\subseteq\mathcal{A}_{f}(v), and

  2. 2.

    {vNG(H)|f(v)|2}\{v\in N_{G}(H)\mid|\mathcal{B}_{f}(v)|\geq 2\} is an independent set in GHG-H.

When there is no ambiguity, we may ignore the subscript ff.

The following lemma states that if we already have a vector assignment of GHG-H, then we can reassign the vectors for vNG(H)v\in N_{G}(H) from (v)\mathcal{B}(v) and the result is also a valid vector assignment.

Lemma 4.1

Let HH be an induced subgraph of a 33-critical graph GG with a bad label π\pi. Let ff be a 33-dim vector assignment on GHG-H with π|GH\pi|_{G-H} and (f(v))vNG(H)(\mathcal{B}_{f}(v))_{v\in N_{G}(H)} an available boundary family. Then for any 33-dim vector assignment gg on GHG-H satisfying

  1. 1.

    g(v)=f(v)g(v)=f(v), vV(GH)NG(H)\forall v\in V(G-H)-N_{G}(H), and

  2. 2.

    g(v)f(v)g(v)\in\mathcal{B}_{f}(v), vNG(H)\forall v\in N_{G}(H),

we have gg is a 33-dim vector assignment on GHG-H with π|GH\pi|_{G-H}.

Proof. We only need to verify that g(v)g(u)=π(uv)g(v)\cdot g(u)=\pi(uv) for all uvE(GH)uv\in E(G-H). Note that A{vV(GH)g(v)f(v)}{vNG(H)|f(v)|2}A\coloneqq\{v\in V(G-H)\mid g(v)\neq f(v)\}\subseteq\{v\in N_{G}(H)\mid|\mathcal{B}_{f}(v)|\geq 2\} from the definition. Then {vV(GH)g(v)f(v)}\{v\in V(G-H)\mid g(v)\neq f(v)\} is an independent set. Since we already have f(v)f(u)=π(uv)f(v)\cdot f(u)=\pi(uv) for all uvE(GH)uv\in E(G-H) and {vV(GH)g(v)f(v)}\{v\in V(G-H)\mid g(v)\neq f(v)\} is an independent set, we now only need to verify that g(v)g(u)=π(uv)g(v)\cdot g(u)=\pi(uv) for all uvE(GH)uv\in E(G-H) satisfying uAu\in A and vAv\notin A. Since g(u)f(u)g(u)\in\mathcal{B}_{f}(u) and g(v)=f(v)g(v)=f(v), we have g(v)g(u)=π(uv)g(v)\cdot g(u)=\pi(uv) by the definition of f(u)\mathcal{B}_{f}(u). \square

We say HH is reducible if there exists an available boundary family (f(v))vNG(H)(\mathcal{B}_{f}(v))_{v\in N_{G}(H)} and a 33-dim vector assignment gg on G[V(H)NG(H)]G[V(H)\cup N_{G}(H)] with π|G[V(H)NG(H)]\pi|_{G[V(H)\cup N_{G}(H)]} such that g(v)f(v)g(v)\in\mathcal{B}_{f}(v) for any vNG(H)v\in N_{G}(H). The following lemma states that there is no reducible structure in 33-critical graph.

Lemma 4.2

Let GG be a 33-critical graph with a bad label π\pi. Then there is no reducible induced subgraph of GG.

Proof. Suppose HH is an induced reducible subgraph of GG. Then GHG-H admits a 33-dim vector assignments ff with π|GH\pi|_{G-H}, an available boundary family (f(v))vNG(H)(\mathcal{B}_{f}(v))_{v\in N_{G}(H)} and a 33-dim vector assignment gg on G[V(H)NG(H)]G[V(H)\cup N_{G}(H)] such that g(v)f(v)g(v)\in\mathcal{B}_{f}(v) for any vNG(H)v\in N_{G}(H). Define a function h:V(G)𝔽23h:V(G)\rightarrow\mathbb{F}_{2}^{3} by letting h(v)=f(v)h(v)=f(v) for any vV(GH)NG(H)v\in V(G-H)-N_{G}(H) and h(v)=g(v)h(v)=g(v) for any vNG(H)V(H)v\in N_{G}(H)\cup V(H). By the definition, h|G[V(H)NG(H)]h|_{G[V(H)\cup N_{G}(H)]} is a 33-dim vector assignment with label π|G[V(H)NG(H)]\pi|_{G[V(H)\cup N_{G}(H)]}. By Lemma 4.1, h|GHh|_{G-H} is a 33-dim vector assignment of GHG-H with label π|GH\pi|_{G-H}. Since there is no edge between V(GH)NG(H)V(G-H)-N_{G}(H) and V(H)V(H), hh is a 33-dim vector assignment of GG with π\pi, a contradiction. \square

In the following, we are going to find certain reducible structures in 33-critical graph.

Lemma 4.3

Let GG be a 33-critical graph with a bad label π\pi. For any vertex vV(G)v\in V(G), at least one edge adjacent to vv is labeled 11 by π\pi.

Proof. Suppose there exists a vertex vV(G)v\in V(G) such that π(uv)=0\pi(uv)=0 for all uNG(v)u\in N_{G}(v). Let G=GvG^{\prime}=G-v. Then GG^{\prime} admits a 33-dim vector assignment ff with π|G\pi|_{G^{\prime}}. Let f(v)=𝟎𝔽23f(v)=\mathbf{0}\in\mathbb{F}_{2}^{3}. Then it is not difficult to verify that ff is a 33-dim vector assignment of GG with π\pi, a contradiction. \square

Lemma 4.4

Let GG be a graph with a label π\pi. If GG admits a 33-dim vector assignment with π\pi, then there exists a 33-dim vector assignment ff with π\pi such that for every vertex vV(G)v\in V(G) of degree at most 22, f(v)𝟎f(v)\neq\mathbf{0}.

Proof. Let ff be the 33-dim vector assignment of GG with π\pi which minimizes nf=|{vV(G)f(v)=𝟎,dG(v)2}|n_{f}=|\{v\in V(G)\mid f(v)=\mathbf{0},d_{G}(v)\leq 2\}|. Suppose nf>0n_{f}>0. Let w{vV(G)f(v)=𝟎,dG(v)2}w\in\{v\in V(G)\mid f(v)=\mathbf{0},d_{G}(v)\leq 2\} and 𝒜(w)={𝐰𝔽23𝐰f(u)=π(uw),uwE(G)}\mathcal{A}(w)=\{\mathbf{w}\in\mathbb{F}_{2}^{3}\mid\mathbf{w\cdot}f(u)=\pi(uw),\forall uw\in E(G)\}. Then |𝒜(w)|2|\mathcal{A}(w)|\geq 2 since dG(w)2d_{G}(w)\leq 2. Choose 𝐰𝒜(w){𝟎}\mathbf{w}\in\mathcal{A}(w)-\{\mathbf{0}\} and define a function g:V(G)𝔽23g:V(G)\rightarrow\mathbb{F}_{2}^{3} by letting g(v)=f(v)g(v)=f(v) for any vV(G){w}v\in V(G)-\{w\} and g(w)=𝐰g(w)=\mathbf{w}. It is easy to verify that gg is a 33-dim vector assignment of GG with π\pi, but ng<nfn_{g}<n_{f}, a contradiction. \square

Lemma 4.5

Let GG be a 33-critical graph of maximum degree 33 with a bad label π\pi. Then GG is 3-regular.

Proof. Suppose there exists a vertex vV(G)v\in V(G) such that d(v)=1d(v)=1. Let uvE(G)uv\in E(G). Then by Lemma 4.3, π(uv)=1\pi(uv)=1. Let V(H)={v}V(H)=\{v\}. Then NG(H)={u}N_{G}(H)=\{u\}. By hypothesis, GvG-v admits a 33-dim vector assignment ff with π|Gv\pi|_{G-v}. Since dGv(u)2d_{G-v}(u)\leq 2, |𝒜f(u)|2|\mathcal{A}_{f}(u)|\geq 2. Let f(u)=𝒜f(u)\mathcal{B}_{f}(u)=\mathcal{A}_{f}(u). Then (f(u))uNG(H)(\mathcal{B}_{f}(u))_{u\in N_{G}(H)} is an available boundary family. Let g(u)(u){𝟎}g(u)\in\mathcal{B}(u)-\{\mathbf{0}\} and we can choose g(v)𝔽23g(v)\in\mathbb{F}_{2}^{3} such that g(v)g(u)=1g(v)\cdot g(u)=1. Then HH is reducible, a contradiction with Lemma 4.2.

Suppose there exists a vertex vV(G)v\in V(G) such that d(v)=2d(v)=2. Let NH(v)={u1,u2}N_{H}(v)=\{u_{1},u_{2}\}. By Lemma 4.3, without loss of generality, assume π(vu1)=1\pi(vu_{1})=1. Let V(H)={v}V(H)=\{v\}. Then NG(H)={u1,u2}N_{G}(H)=\{u_{1},u_{2}\}. By hypothesis and Lemma 4.4, GvG-v admits a 33-dim vector assignment ff with π|Gv\pi|_{G-v} such that f(u1),f(u2)𝟎f(u_{1}),f(u_{2})\neq\mathbf{0}. Let (u1)={f(u1)}\mathcal{B}(u_{1})=\{f(u_{1})\} and (u2)=𝒜(u2)\mathcal{B}(u_{2})=\mathcal{A}(u_{2}). Then ((ui))i=1,2(\mathcal{B}(u_{i}))_{i=1,2} is an available boundary family. Since dGv(u2)2d_{G-v}(u_{2})\leq 2, we have |(u2)|2|\mathcal{B}(u_{2})|\geq 2. Let g(u1)=f(u1)g(u_{1})=f(u_{1}). If π(vu2)=1\pi(vu_{2})=1 (resp. π(vu2)=0\pi(vu_{2})=0), choose g(u2)(u2){𝟎}g(u_{2})\in\mathcal{B}(u_{2})-\{\mathbf{0}\} (resp. g(u2)(u2){f(u1)}g(u_{2})\in\mathcal{B}(u_{2})-\{f(u_{1})\}). It is easy to verify in either case, there exists g(v)𝔽23g(v)\in\mathbb{F}_{2}^{3} such that g(v)g(ui)=π(vui)g(v)\cdot g(u_{i})=\pi(vu_{i}) for i=1,2i=1,2. So HH is reducible, a contradiction with Lemma 4.2. \square

Refer to caption

Figure 2: K4K_{4}^{-} in GG.

Refer to caption

Figure 3: Triangle in GG.
Lemma 4.6

Let GG be a 33-critical 3-regular graph with a bad label π\pi. There is no induced K4K_{4}^{-} in GG, where K4K_{4}^{-} is the graph obtained by deleting an edge in K4K_{4}.

Proof. Suppose there exists a K4K_{4}^{-} in GG with vertex set {v0,v1,u0,u1}\{v_{0},v_{1},u_{0},u_{1}\} and u0u1E(G)u_{0}u_{1}\notin E(G) (See Figure 3).

Let H=G[{v0,v1}]H=G[\{v_{0},v_{1}\}]. Then NG(H)={u0,u1}N_{G}(H)=\{u_{0},u_{1}\}. By hypothesis and Lemma 4.4, GHG-H admits a 33-dim vector assignment ff with π|GH\pi|_{G-H} such that f(u0),f(u1)𝟎f(u_{0}),f(u_{1})\neq\mathbf{0}. Let (ui)=𝒜f(ui),i=0,1\mathcal{B}(u_{i})=\mathcal{A}_{f}(u_{i}),i=0,1. Then ((ui))i=0,1(\mathcal{B}(u_{i}))_{i=0,1} is an available boundary family. We have the following properties:

  1. 1.

    For each i{0,1}i\in\{0,1\}, |(ui)|4|\mathcal{B}(u_{i})|\geq 4 by dGH(ui)=1d_{G-H}(u_{i})=1.

  2. 2.

    For each i{0,1}i\in\{0,1\}, if π(v0ui)=π(v1ui)=0\pi(v_{0}u_{i})=\pi(v_{1}u_{i})=0, then 𝟎(ui)\mathbf{0}\notin\mathcal{B}(u_{i}) by Lemma 4.3.

  3. 3.

    For each i{0,1}i\in\{0,1\}, at least one edge in {v0v1,viu0,viu1}\{v_{0}v_{1},v_{i}u_{0},v_{i}u_{1}\} is labeled one by Lemma 4.3.

With above properties, we claim that HH is reducible. The claim is proved by computer by enumerating all available boundary families with above properties. The source codes can be found on GitHub. Therefore, we derive a contradiction with Lemma 4.2. \square

Lemma 4.7

Let GG be a 33-critical 3-regular graph with a bad label π\pi. Then there is no triangle in GG.

Proof. Suppose there exists a triangle with vertices {v0,v1,v2}\{v_{0},v_{1},v_{2}\} and let uiu_{i} be the neighbor of viv_{i}, for i=0,1,2i=0,1,2 (See Figure 3). By Lemma 4.6, u0,u1,u2u_{0},u_{1},u_{2} are either distinct vertices, or u0=u1=u2u_{0}=u_{1}=u_{2}. If u0=u1=u2u_{0}=u_{1}=u_{2}, then G=K4G=K_{4} by GG being 3-regular. However, diam((K4))=3diam(\mathcal{I}(K_{4}))=3 [5], which contradicts that GG is 33-critical. Hence, we conclude that u0,u1,u2u_{0},u_{1},u_{2} are distinct vertices. Let V(H)={v0,v1,v2}V(H)=\{v_{0},v_{1},v_{2}\}. Then NG(H)={u0,u1,u2}N_{G}(H)=\{u_{0},u_{1},u_{2}\}. By hypothesis and Lemma 4.4, GHG-H admits a 33-dim vector assignment ff with π|GH\pi|_{G-H} such that f(ui)𝟎,i=0,1,2f(u_{i})\neq\mathbf{0},i=0,1,2. We can assume, without loss of generality, that u0u_{0} satisfies the property: if f(u1)=f(u2)f(u_{1})=f(u_{2}), then f(u0)=f(u1)=f(u2)f(u_{0})=f(u_{1})=f(u_{2}). Let (u0)=𝒜(u0)\mathcal{B}(u_{0})=\mathcal{A}(u_{0}) and (ui)={f(ui)},i=1,2\mathcal{B}(u_{i})=\{f(u_{i})\},i=1,2. Now we have the following properties:

  1. 1.

    |(u0)|2|\mathcal{B}(u_{0})|\geq 2 by dGH(u0)=2d_{G-H}(u_{0})=2.

  2. 2.

    For each i=0,1,2i=0,1,2, at least one edge adjacent to viv_{i} is labeled one by Lemma 4.3.

  3. 3.

    If π(u0v0)=0\pi(u_{0}v_{0})=0, then 𝟎(u0)\mathbf{0}\notin\mathcal{B}(u_{0}), also by Lemma 4.3.

  4. 4.

    If f(u1)=f(u2)f(u_{1})=f(u_{2}), then f(u1)=f(u0)(u0)f(u_{1})=f(u_{0})\in\mathcal{B}(u_{0}).

With above properties, we claim that HH is reducible which is proved by computer. The source codes can be found on GitHub. Therefore, we derive a contradiction with Lemma 4.2. \square

Refer to caption

Figure 4: P3P_{3} with edges labeled one in GG.

Refer to caption

Figure 5: K2,3K_{2,3} in GG.
Lemma 4.8

Let GG be a 33-critical 3-regular graph with a bad label π\pi. Then there is no P3P_{3} with two edges labeled one in GG.

Proof. Suppose there exists a path wu0uwu_{0}u^{\prime} such that π(wu0)=π(u0u)=1\pi(wu_{0})=\pi(u_{0}u^{\prime})=1. By Lemma 4.7, uwE(G)u^{\prime}w\notin E(G). Let u1,u2u_{1},u_{2} be the neighbor of ww (See Figure 5). By Lemma 4.7, {u0,u1,u2}\{u_{0},u_{1},u_{2}\} is an independent set. Let V(H)={w}V(H)=\{w\}. Then NG(H)={u0,u1,u2}N_{G}(H)=\{u_{0},u_{1},u_{2}\}. By hypothesis, GHG-H admits a 33-dim vector assignment ff with π|GH\pi|_{G-H}. Let (ui)=𝒜(ui),i=0,1,2\mathcal{B}(u_{i})=\mathcal{A}(u_{i}),i=0,1,2. Then ((ui))i=0,1,2(\mathcal{B}(u_{i}))_{i=0,1,2} is an available boundary family. We have the following properties:

  1. 1.

    For each i{0,1,2}i\in\{0,1,2\}, |(ui)|2|\mathcal{B}(u_{i})|\geq 2 by dGH(ui)=2d_{G-H}(u_{i})=2.

  2. 2.

    𝟎(u0)\mathbf{0}\notin\mathcal{B}(u_{0}), because π(u0u)=1\pi(u_{0}u^{\prime})=1.

  3. 3.

    For each i=1,2i=1,2, if π(wui)=0\pi(wu_{i})=0, then 𝟎(ui)\mathbf{0}\notin\mathcal{B}(u_{i}) by Lemma 4.3.

With above properties, we claim that HH is reducible which is proved by computer (GitHub). Therefore, we derive a contradiction with Lemma 4.2. \square

Lemma 4.9

Let GG be a 33-critical 3-regular graph with a bad label π\pi. Then there is no K2,3K_{2,3} in GG.

Proof. Suppose there exists a K2,3K_{2,3} with vertices {vi}i=0,1{ui}i=0,1,2\{v_{i}\}_{i=0,1}\cup\{u_{i}\}_{i=0,1,2} and uivjE(G)u_{i}v_{j}\in E(G) for every i=0,1i=0,1 and j=0,1,2j=0,1,2 (See Figure 5). By Lemmas 4.3 and 4.8, without loss of generality, we can assume π(v0u0)=π(v1u2)=1\pi(v_{0}u_{0})=\pi(v_{1}u_{2})=1 and other edges in K2,3K_{2,3} are labeled zero. By Lemma 4.7, {ui}i=0,1,2\{u_{i}\}_{i=0,1,2} is an independent set. Let H=G[{v0,v1}]H=G[\{v_{0},v_{1}\}]. Then NG(H)={u0,u1,u2}N_{G}(H)=\{u_{0},u_{1},u_{2}\}. By hypothesis, GHG-H admits a 33-dim vector assignment ff. Let (ui)=𝒜(ui),i=0,1,2\mathcal{B}(u_{i})=\mathcal{A}(u_{i}),i=0,1,2. Then ((ui))i=0,1,2(\mathcal{B}(u_{i}))_{i=0,1,2} is an available boundary family. We have the following properties:

  1. 1.

    For each i{0,1,2}i\in\{0,1,2\}, |(ui)|4|\mathcal{B}(u_{i})|\geq 4 by dGH(ui)=1d_{G-H}(u_{i})=1.

  2. 2.

    By Lemma 4.3, 𝟎(u1)\mathbf{0}\notin\mathcal{B}(u_{1}) and 𝟎(ui)\mathbf{0}\in\mathcal{B}(u_{i}) for i=0,2i=0,2.

With above properties, we claim that HH is reducible which is proved by computer (GitHub). Therefore, we derive a contradiction with Lemma 4.2. \square

Refer to caption

Figure 6: C4C_{4} with at least one edge labeled one in GG.

Refer to caption

Figure 7: Edge labeled one in GG.
Lemma 4.10

Let GG be a 33-critical 3-regular graph with a bad label π\pi. Then there is no C4C_{4} with at least on edge is labeled one in GG.

Proof. Suppose there exists a C4C_{4} with vertices {vi}i=0,1,2,3\{v_{i}\}_{i=0,1,2,3} and π(v0v1)=1\pi(v_{0}v_{1})=1. Let uiu_{i} be the neighbor of viv_{i} for i=0,1,2,3i=0,1,2,3 (See Figure 7). By Lemmas 4.7 and 4.9, {ui}i=0,1,2,3\{u_{i}\}_{i=0,1,2,3} are distinct vertices. By Lemma 4.8, π(v0v2)=π(v1v3)=0\pi(v_{0}v_{2})=\pi(v_{1}v_{3})=0. Let H=C4H=C_{4}. Then NG(H)={u0,u1,u2,u3}N_{G}(H)=\{u_{0},u_{1},u_{2},u_{3}\}. By hypothesis and Lemma 4.4, GHG-H admits a 33-dim vector assignment ff with π|GH\pi|_{G-H} such that f(ui)𝟎,i=0,1,2,3f(u_{i})\neq\mathbf{0},i=0,1,2,3.

Let us first consider the case π(v2v3)=0\pi(v_{2}v_{3})=0. In this case, let (ui)={f(ui)},i=0,1,2,3\mathcal{B}(u_{i})=\{f(u_{i})\},i=0,1,2,3. Then ((ui))i=0,1,2,3(\mathcal{B}(u_{i}))_{i=0,1,2,3} is an available boundary family. We claim HH is reducible with ((ui))i=0,1,2,3(\mathcal{B}(u_{i}))_{i=0,1,2,3} which is proved by computer (GitHub), a contradiction with Lemma 4.2.

Now suppose π(v2v3)=1\pi(v_{2}v_{3})=1. Then π(viui)=0,i=0,1,2,3\pi(v_{i}u_{i})=0,i=0,1,2,3 by Lemma 4.8. Since GG is 3-regular, there exists t{1,2,3}t\in\{1,2,3\} such that u0utE(G)u_{0}u_{t}\notin E(G). Let (ui)=𝒜(ui)\mathcal{B}(u_{i})=\mathcal{A}(u_{i}) for i=0,ti=0,t and (ui)={f(ui)}\mathcal{B}(u_{i})=\{f(u_{i})\} for other ii. Then ((ui))i=0,1,2,3(\mathcal{B}(u_{i}))_{i=0,1,2,3} is an available boundary family. Since π(viui)=0\pi(v_{i}u_{i})=0 for each ii, we have 𝟎(ui)\mathbf{0}\notin\mathcal{B}(u_{i}) by Lemma 4.3. Moreover, for i=0,ti=0,t, we have |(ui)|2|\mathcal{B}(u_{i})|\geq 2 since dGH(ui)=2d_{G-H}(u_{i})=2. We claim that we can choose 𝐮𝐢(ui)\mathbf{u_{i}}\in\mathcal{B}(u_{i}) for any i{0,1,2,3}i\in\{0,1,2,3\} such that 𝐮𝐣𝟎=𝐮𝐣𝟏\mathbf{u_{j_{0}}=u_{j_{1}}} and 𝐮𝐣𝟐=𝐮𝐣𝟑\mathbf{u_{j_{2}}=u_{j_{3}}} do not occur, where {j0,j1,j2,j3}={0,1,2,3}\{j_{0},j_{1},j_{2},j_{3}\}=\{0,1,2,3\}. The claim can be proved by simple discussion. Then define g:V(GH)𝔽23g:V(G-H)\rightarrow\mathbb{F}_{2}^{3} by letting g(ui)=𝐮𝐢,i=0,1,2,3g(u_{i})=\mathbf{u_{i}},i=0,1,2,3 and g(v)=f(v)g(v)=f(v) for all other vertex vv. By lemma 4.1, gg is a 33-dim vector assignment of GHG-H with π|GH\pi|_{G-H}.

Let g(ui)=g(ui),i=0,1,2,3\mathcal{B}_{g}(u_{i})=g(u_{i}),i=0,1,2,3, then (g(ui))i=0,1,2,3(\mathcal{B}_{g}(u_{i}))_{i=0,1,2,3} is an available boundary family. We claim HH is reducible with (g(ui))i=0,1,2,3(\mathcal{B}_{g}(u_{i}))_{i=0,1,2,3} which is proved by computer (GitHub). A contradiction. \square

Now we have plenty of forbidden structures in GG, and we can finally prove Theorem 1.4.

Proof of Theorem 1.4. By contradiction. Let GG be the counter example with minimum number of vertices and then minimum edges. Then GG is 33-critical. Since Δ(G)3\Delta(G)\leq 3, GG is 3-regular by Lemma 4.5. Let π\pi be a bad label of GG.

We assume v0v1E(G)v_{0}v_{1}\in E(G) satisfying π(v0v1)=1\pi(v_{0}v_{1})=1 by Lemma 4.3. Let u0,u1u_{0},u_{1} be the neighbor of v0v_{0} and u2,u3u_{2},u_{3} the neighbor of v1v_{1} (See Figure 7). By Lemmas 4.7 and 4.10, {ui}1,2,3,4\{u_{i}\}_{1,2,3,4} is an independent set. Let H=G[{v0,v1}]H=G[\{v_{0},v_{1}\}]. Then NG(H)={u0,u1,u2,u3}N_{G}(H)=\{u_{0},u_{1},u_{2},u_{3}\}. By hypothesis, GHG-H admits a 33-dim vector assignment ff. Let (ui)=𝒜(ui),i=0,1,2,3\mathcal{B}(u_{i})=\mathcal{A}(u_{i}),i=0,1,2,3. Then ((ui))i=0,1,2,3(\mathcal{B}(u_{i}))_{i=0,1,2,3} is an available boundary family. Since dGH(ui)=2,i=0,1,2,3d_{G-H}(u_{i})=2,i=0,1,2,3, we have |(ui)|2,i=0,1,2,3|\mathcal{B}(u_{i})|\geq 2,i=0,1,2,3. By Lemma 4.4, we have 𝟎(ui)\mathbf{0}\notin\mathcal{B}(u_{i}) for every i=0,1,2,3i=0,1,2,3. Then we claim that HH is reducible which is proved by computer (GitHub), a contradiction with Lemma 4.2. \square

Acknowledgement

Y. Yang is supported by the Fundamental Research Funds for the Central University (Grant 500423306) in China. M. Lu is supported by the National Natural Science Foundation of China (Grant 12171272 & 12161141003).

References

  • [1] N. Alon, E. Powierski, M. Savery, A. Scott, and E. Wilmer. Invertibility of digraphs and tournaments. SIAM Journal on Discrete Mathematics, 38(1):327–347, 2024.
  • [2] G. Aubian, F. Havet, F. Hörsch, F. Klingelhoefer, N. Nisse, C. Rambaud, and Q. Vermande. Problems, proofs, and disproofs on the inversion number. arXiv preprint arXiv:2212.09188, 2022.
  • [3] J. Bang-Jensen, J. C. F. da Silva, and F. Havet. On the inversion number of oriented graphs. Discrete Mathematics & Theoretical Computer Science, 23(Special issues), 2022.
  • [4] H. Belkhechine, M. Bouaziz, I. Boudabbous, and M. Pouzet. Inversion dans les tournois. Comptes Rendus. Mathématique, 348(13-14):703–707, 2010.
  • [5] F. Havet, F. Hörsch, and C. Rembaud. Diameter of the inversion graph. arXiv preprint arXiv:2405.04119, 2024.
  • [6] P. Scheffler. Die Baumweite von Graphen als ein Maß für die Kompliziertheit algorithmischer Probleme, volume 4. Akademie der Wissenschaften der DDR, Karl-Weierstrass-Institut für Mathematik, 1989.
  • [7] J. van Leeuwen. Graph algorithms. In Algorithms and complexity, pages 525–631. Elsevier, 1990.